are breaking waves, bores, surges and jumps the same...

31
ORIGINAL ARTICLE Are breaking waves, bores, surges and jumps the same flow? Pierre Lubin 1,2 Hubert Chanson 2 Received: 26 July 2015 / Accepted: 21 July 2016 / Published online: 1 August 2016 Ó Springer Science+Business Media Dordrecht 2016 Abstract The flow structure in the aerated region of the roller generated by breaking waves remains a great challenge to study, with large quantities of entrained air and tur- bulence interactions making it very difficult to investigate in details. A number of analogies were proposed between breaking waves in deep or shallow water, tidal bores and hydraulic jumps. Many numerical models used to simulate waves in the surf zone do not implicitly simulate the breaking process of the waves, but are required to parameterise the wave-breaking effects, thus relying on experimental data. Analogies are also assumed to quantify the roller dynamics and the energy dissipation. The scope of this paper is to review the different analogies proposed in the literature and to discuss current practices. A thorough survey is offered and a discussion is developed an aimed at improving the use of possible breaking proxies. The most recent data are revisited and scrutinised for the use of most advanced numerical models to educe the surf zone hydrodynamics. In particular, the roller dynamics and geometrical characteristics are discussed. An open discussion is proposed to explore the actual practices and propose perspectives based on the most appropriate analogy, namely the tidal bore. Keywords Breaking waves Breaking bores Hydraulic jumps Air bubble entrainment Flow singularity Tidal bores 1 Introduction Surface wave breaking, occurring in the open ocean or the coastal zone, is a complex and challenging two-phase flow phenomenon which plays an important role in numerous processes, including air–sea transfer of gas, momentum and energy, and in a number of & Pierre Lubin [email protected] 1 Universite ´ de Bordeaux, I2M, CNRS UMR 5295, 16 Avenue Pey-Berland, 33607 Pessac, France 2 School of Civil Engineering, The University of Queensland, Brisbane, QLD 4072, Australia 123 Environ Fluid Mech (2017) 17:47–77 DOI 10.1007/s10652-016-9475-y

Upload: others

Post on 22-Sep-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

ORIGINAL ARTICLE

Are breaking waves, bores, surges and jumps the sameflow?

Pierre Lubin1,2 • Hubert Chanson2

Received: 26 July 2015 / Accepted: 21 July 2016 / Published online: 1 August 2016� Springer Science+Business Media Dordrecht 2016

Abstract The flow structure in the aerated region of the roller generated by breaking

waves remains a great challenge to study, with large quantities of entrained air and tur-

bulence interactions making it very difficult to investigate in details. A number of

analogies were proposed between breaking waves in deep or shallow water, tidal bores and

hydraulic jumps. Many numerical models used to simulate waves in the surf zone do not

implicitly simulate the breaking process of the waves, but are required to parameterise the

wave-breaking effects, thus relying on experimental data. Analogies are also assumed to

quantify the roller dynamics and the energy dissipation. The scope of this paper is to

review the different analogies proposed in the literature and to discuss current practices. A

thorough survey is offered and a discussion is developed an aimed at improving the use of

possible breaking proxies. The most recent data are revisited and scrutinised for the use of

most advanced numerical models to educe the surf zone hydrodynamics. In particular, the

roller dynamics and geometrical characteristics are discussed. An open discussion is

proposed to explore the actual practices and propose perspectives based on the most

appropriate analogy, namely the tidal bore.

Keywords Breaking waves � Breaking bores � Hydraulic jumps � Air bubble entrainment �Flow singularity � Tidal bores

1 Introduction

Surface wave breaking, occurring in the open ocean or the coastal zone, is a complex and

challenging two-phase flow phenomenon which plays an important role in numerous

processes, including air–sea transfer of gas, momentum and energy, and in a number of

& Pierre [email protected]

1 Universite de Bordeaux, I2M, CNRS UMR 5295, 16 Avenue Pey-Berland, 33607 Pessac, France

2 School of Civil Engineering, The University of Queensland, Brisbane, QLD 4072, Australia

123

Environ Fluid Mech (2017) 17:47–77DOI 10.1007/s10652-016-9475-y

Page 2: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

technical applications such as acoustic underwater communications and optical properties

of the water column. The major visible feature during wave breaking is the large quantities

of air entrained in the form of bubble clouds and whitecaps, generally coined surface foam

(Figs. 1, 2). The generation of bubble clouds has been shown to induce energy dissipation

and turbulent mixing, to contribute to heat exchange and enhance gas transfer [94, 182].

Bubble clouds have been shown to influence climate and intensification of tropical

cyclones [176], and cause the ocean ambient noise [148]. The breakup and evolution of

entrained air into numerous bubbles is a source of acoustic noise, which is important for

naval hydrodynamics. The hydrodynamic performance of ships is influenced by the wake

Fig. 1 Sketches of breaking waves, bore, roller and hydraulic jump

48 Environ Fluid Mech (2017) 17:47–77

123

Page 3: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

modified by the air entrainment, and the sound generated by the bubble clouds render the

ships subject to detection. In hydraulic engineering, large spillways are often protected

from cavitation damage by controlling aeration [76, 159].

(A)

(B)

(A1)

(A2)

(B1) (B2)

(C)

Fig. 2 Air entrainment in breaking waves, bores and jumps. a Breaking waves. a1 Plunging breaking offMain Beach, North Stradbroke Island (Australia) on 19 April 2014—Shutter speed: 1/400 s, a2 spillingbreaker off Pointe du Grand Minou (Brest, France) on 18 March 2004—Shutter speed: 1/800 s. b Breakingtidal bore of the Qiantang River (China) on 11 October 2014. b1 Bore approaching the old seawall atXinchang at 12:50, b2 Bore roller propagating (from left to right) parallel to the seawall at Qilimiao at13:20—The roller height was about 2.5 to 3 m, the bore celerity 4 to 5 m/s and Fr *2—Shutter speed:1/6400 s. c Hydraulic jump roller (steady flow direction from right to left): Fr = 7.5, Re = 1.4 9 105,d1 = 0.03 m, shutter speed: 1/1000 s

Environ Fluid Mech (2017) 17:47–77 49

123

Page 4: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

Many numerical models (e.g. Boussinesq equations) used to study waves in the surf

zone do not implicitly simulate the breaking process of the waves [54]. The wave-breaking

effects have to be parameterised by incorporating additional terms in the mass and

momentum equations (e.g. [13, 55, 100, 141, 171]. The challenge is to take the breaking

process into account to ensure an accurate description of the surf zone, including the wave

height decay and the setup development. The main consideration is to dissipate energy

when wave breaking is likely to occur. Svendsen [167, 168] proposed the roller concept, in

the form of a volume of water carried shoreward with the wave. Local roller thickness and

mean front slope of the breaker were used to quantify part of the local momentum deficit.

But the vertical surface roller of the breaking wave is only considered to play an important

part in the momentum and energy conservation. However, the energy flux and dissipation

during wave breaking remain difficult to quantify. Most recent modelling attempts are still

struggling with the lack of physical knowledge of the finest details of the breaking pro-

cesses, which makes the task of parameterising breaking effects very difficult since no

universal scaling laws for physical variables have been proposed so far. Physical param-

eters, such as the height and length of the roller, have to be quantified and criteria have to

be defined with critical bounded values to estimate where the waves break and stop

breaking. Thus models still need calibration and further improvements [22].

The turbulent flow dynamics in bubble clouds is a very challenging numerical problem.

Esmaeeli and Tryggvason [75] studied direct numerical simulations of buoyant bubbles in

a two-dimensional periodic domain. They simulated 144 and 324 bubbles, showing that the

work done by the buoyant bubbles increased the energy of flow structures much larger than

the bubbles. But 3D direct modeling of air bubble entrainment and evolution at the scale of

the surf zone is computationally unaffordable. Another way of tackling dispersed two-

phase flows is using a continuum-mechanical approach [71]. Two-fluid models are used to

model the polydisperse two-fluid bubbly flow based on mixture theory

[26, 68, 127, 134, 162]. A first attempt to use a continuum type model for studying bubbly

flow under surface breaking waves was made by Shi et al. [162]. They proposed a phys-

ically-based numerical model for prediction of air bubble population in a surf zone-scale

domain. The air entrainment was formulated by connecting the shear production at air–

water interface and the bubble number density with the bubble size spectra as observed by

Deane and Stokes [67]. The model was initially fed with the entrained bubbles and used to

simulate the evolution of the bubble plumes. This approach requires much less spatial and

temporal resolution than needed to capture detailed air entrainment process in DNS sim-

ulations. The model results revealed that bubbles larger than 1 mm provide a major

contribution to void fraction, while smaller bubbles contribute significantly to the cumu-

lative interfacial area of the bubble cloud but do not contribute much to the total volume of

air. Discrepancies between observations and model behaviour were nevertheless reported.

Based on the works of Ma et al. [127], Derakhti and Kirby [68] used an Eulerian–Eulerian

polydisperse two-fluid model in an LES framework. Detailed overviews on methods and

models for CFD of multiphase flows can be found in textbooks [61, 72]. More information

about turbulence modelling in the framework of multiphase flows is given by Labourasse

et al. [107] and Bombardelli [17]. Smoothed-particle hydrodynamics (SPH) is also a mesh-

free method which can be used to describe accurately the 3D surf zone hydrodynamics, as

recently shown by Farahani and Dalrymple [77] who investigated some novel coherent

turbulent vortical structures under broken solitary waves. The state-of-the-art is detailed by

Gomez-Gesteira et al. [84] and Violeau and Rogers [177], who detailed a number of

examples in which SPH simulations have been successfully used in fluid flow research and

hydraulic engineering.

50 Environ Fluid Mech (2017) 17:47–77

123

Page 5: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

Numerical models still rely on experimental data. Detailed information on the

temporal and spatial variations of the void fraction fields beneath breaking waves is

required. Instantaneous void fraction and interfacial velocity data are critically needed

to calibrate and improve numerical models of the two-phase flow generated beneath

plunging and spilling breaking waves. Models for air-entrainment are critically

dependent upon accurate estimates of the surface area affected by wave breaking.

Controlled laboratory experiments and accurate measurements of void fraction and

bubble size distributions beneath plunging and spilling breakers are still very chal-

lenging. When a wave breaks, the tip forms a liquid jet which impinges on the front

face of the wave and creates an air cavity which breaks into bubbles. The character-

isation of the bubble sizes resulting from the cavity collapse has to be measured and

the trajectories of these entrained bubbles are also critical information. The initial

stages of the breaking of a wave generated a large amount of bubbles production and

to the distribution at greater depths. The bubble clouds will then form, grow and decay

during the propagation of the turbulent air/water mixing region forming the bore, the

temporal variations of all bubble cloud dimensions reflecting this evolution. The large

volumes of air in bubbles rapidly evolve into a distribution of bubble sizes which

interacts with liquid turbulence and organised motions.

The motion of bubbles relative to the liquid causes velocity fluctuations in the

water column and increases the energy of liquid motion at the scales comparable with

the bubble diameter [68]. Bubble plume kinematics and dynamics, and the structure of

the turbulent bubbly flow under breaking waves constitute critical information to be

taken into account for an accurate description of the wave breaking process [129].

While the former can be studied experimentally, the liquid–bubble interactions, i.e. the

effects of dispersed bubbles on organised and turbulent motions, are still poorly

understood.

When looking at a bore, whereas it has been generated by a stationary hydraulic jump, a

surface wave breaking on the ocean or in the surf zone, or a tidal bore propagating

upstream a river, the question is: are we looking at the same flow? Is there only one bore

structure, or are there variations depending on the initial conditions leading to its occur-

rence and behaviour? To what extent can we compare the bores and use the quantities

through similarity? It is the aim of this contribution to contribute to the transfer of

knowledge from detailed measurements realised in hydraulic jumps and tidal bores, to the

wave breaking investigation. The first part of the article is dedicated to the identification of

the knowledge gaps encountered when attempting to simulate numerically the hydrody-

namics of breaking waves and a review of the various analogies which have been proposed

in the literature. The next part reports on the state-of-the-art of the studies focusing on the

void fraction and velocity analysis under breaking waves, tidal bores and hydraulic jumps.

Based on this survey, we attempt to identify and assess the quantities which can be

considered for possible analogies. The most recent data are revisited and scrutinised for the

use of most advanced numerical models to educe the surf zone hydrodynamics, highly

influenced by the wave breaking process. An open discussion is proposed to explore the

actual practises and propose perspectives based on the most appropriate analogy, namely

the tidal bore.

Environ Fluid Mech (2017) 17:47–77 51

123

Page 6: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

2 Knowldege gaps for the modelling of the surf zone hydrodynamics

2.1 Current state of practice in numerical modelling and limitations

Most numerical models only consider macro-scale roller properties. The roller formation

and propagation have been shown to be a highly unsteady process, with air entrainment and

turbulence generation. The most advanced models, which are generally used to simulate

non-linear wave transformations in coastal areas, are based either on the non-linear shallow

water equations (NSW), the Boussinesq-type equations (BT), or some form of hybrid

model. Extensive developments and break-through progress have been made recently for a

large variety of coastal engineering applications (e.g. [3, 22, 100, 171]). A key feature, the

breaking process, is however not explicitly simulated and missing in these models. Several

approaches and parametrisations have thus been proposed to introduce wave breaking in

NSW and BT models. Any such approach requires the quantification of energy dissipation,

dynamically activated when wave breaking is likely to occur. Some physically based

criteria have to be able to activate or deactivate these extra terms and simple expressions

are generally favoured. Simple quantities include geometrical aspects of the roller,

including heights, lengths and angles, easily extracted from any visual observations in

laboratory and in the field, All these quantities cannot be estimated from a single exper-

iment. Instead a composite set of data and practices have been elaborated though time by

looking at various analogue flows, and some variations have been proposed in order to fill

the gaps.

Practically, most numerical models need to evaluate:

1. A Froude number Fr characteristic of wave breaking, of when it occurs and stops (with

Fr varying with water depth). Currently, an accepted value for the transition between

non breaking and breaking waves has been identified in Froude number range between

1.3 and 1.6 [145]. This is based upon the analogy with non-breaking undular hydraulic

jump and bore [46, 78, 104, 117, 173];

2. The roller height hr, derived from momentum considerations (see Appendix 2);

3. The roller length Lr, determined empirically. A common parameterisation is

Lr = 2.91 9 hr [87], although the re-analysis of large-scale experiments suggests

Lr/hr & 1 to 8 (Fig. 5). In Fig. 5, steady breaker, stationary hydraulic jump and tidal

bore data are compared;

4. The mean front slope angle / [161], typically between 8� and 30� for the termination

and initiation of the breaking event respectively;

5. The roller celerity (or celerity of the breaking wave);

6. The energy dissipation in the roller region;

7. The bubble size distributions, often improperly estimated based upon Hinze’s [89]

model developed in the case of a single droplet under non-coalescing conditions (!).

To estimate most of these quantities, flow analogies have been considered, but some

limitations are clearly identified and some modifications, based on new experimental data

analysis, are proposed in the following sections.

2.2 Flow analogies or not?

A number of analogies were proposed between breaking waves, bores and jumps (Ap-

pendix 1). Appendix 1 lists a number of early seminal references and Fig. 1 presents

52 Environ Fluid Mech (2017) 17:47–77

123

Page 7: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

definition sketches. The steady breaker configuration was proposed as a simplification of

the spilling breaker [6, 7]. Important results were obtained [8, 59, 63, 73, 121], but there is

still on-going argument about the validity of this analogy [102]. Further links were

developed between breaking waves and steady flow configurations. These encompassed

comparisons between plunging breakers and plunging jets [44, 47, 48, 56, 93, 143, 160],

between spilling breakers and stationary hydraulic jumps [23, 24, 123, 128, 147] and

between spilling breakers and translating hydraulic jumps (also called positive surges or

tidal bores) [24, 123, 147]. In parallel, there have been numerous discussions about the

similarities and differences between stationary and translating hydraulic jumps (e.g.

[65, 118, 164, 174], although the open channel hydraulic literature develops the same

integral approach for both types [32, 41, 88, 118].

To date, the mechanistic connections between these flows are not well understood and

have not always been successful. Wave-plunging jet conditions appear to produce a

qualitatively different type of impact, with almost no penetration into the oncoming flow

and a pronounced splash that cascades multiple times down the face of the wave. What is

better characterised however, is the volume of air trapped by the initial contact of the jet

with the wave face, which has been numerically simulated, and its shape has been suc-

cessfully modelled, at least for a limited set of conditions (e.g. [126]. However, the

processes that follow the initial contact are only known qualitatively for the majority of the

breaking conditions, and thus still require further study in order to acquire improved

physical understanding. Furthermore, wave breaking is a combination of transient pro-

cesses which evolve within the breaking duration making adequate physical understanding

a challenging proposition. Overall, ‘‘In studying any turbulent flow it is very helpful if it

can be shown to be similar to other well known flows’’ [147]. Below, a number of seminal

flow configurations are explored and the relevance of flow comparisons is discussed.

3 Void fraction kinematics

3.1 Breaking waves

Fuhrboter [79] discussed the correlation between the turbulence generated in the surf zone

and the amount of air entrained during the breaking of the waves, as well as the sudden

reduction of wave height and energy. He highlighted the importance to study quantitatively

the air entrainment process for a detailed comprehension of the surf zone physics. Vagle

and Farmer [175] and recently Anguelova and Huq [1] reviewed the different techniques

used to quantify the void fraction under breaking waves. Both works concluded that

combined techniques were the best approach. Indeed, the higher the concentrations of

bubbles within bubble clouds, the more difficult it is to count and measure individual

bubbles.

Some studies have been conducted in field while others have been completed in physical

wave tanks. Thorpe [170] studied wind-waves breaking and speculated that wind speed,

salinity, and temperature were major factors, possibly responsible for existing discrepan-

cies that arised when comparing data from different sources. Monahan [133] proposed the

terms Alpha-plume (high void fraction, short lifespan), Beta-plume, and Gamma-plume

(low void fraction, long lifespan) to describe the evolution of a bubble cloud, from its

formation to its disappearance (e.g. dissolution, degassing and advection). Most field

studies confirmed that the Alpha-plumes consist of high void fractions (10 % or more) with

Environ Fluid Mech (2017) 17:47–77 53

123

Page 8: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

large bubble sizes (radii ranging from tens of micrometers to millimeters). At the other end

of the process, the Gamma-plume were observed to be very low void fraction between

10-5 and 10-8 and containing bubbles with radii on the order of O(10–100) lm. The

lifetime of a whole bubble cloud may be about a hundred of seconds. The bubble clouds

are also generally confined to the first few meters of the water column. For example,

Lamarre and Melville [110] compared field and laboratory void fraction measurements

obtained with an impedance probe, and showed large void fraction values at shallow

locations while lower void fraction values were found deeper. Deane [66] used acoustic

and optical measurements of individual breaking waves in the surf zone, off La Jolla

Shores beach, California. Total void fractions of 0.3–0.4 were measured, consisting of

bubbles with radius greater than 1 mm. Stokes and Dean [165] observed that the time scale

for the generation of clouds of submillimetric bubbles was on the order of about 90 ms.

Dahl and Jessup [64] found comparable quantities in deep-ocean studies. Gemmrich and

Farmer [83] measured void fraction values (e.g. 10-2 at 0.25 m below the free-surface),

associated with low penetrating breaking events (spilling breakers), while they speculated

that higher values of void fractions found deeper would be associated with more energetic

violent events (plunging breakers). Interestingly Gemmrich [82] found higher turbulence

levels within the wave crest region of the breaking waves, suggesting that the bubble

fragmentation process is mainly driven by turbulence. Most studies reported that void

fraction changes significantly during the lifetime of the bubble cloud, from high void

fractions in the first seconds of the breaking event to residual void fractions persisting for

long times. Most field studies consisted in wind-waves breaking observations, with only

few events giving data susceptible to be accurately analysed.

A lot of studies investigated the hydrodynamics in the surf zone, especially the general

mechanisms involved during the breaking process [146], the generation of turbulence [10],

and sediment transport. When waves break, the flow suddenly exhibits a violent transition

from irrotational to rotational motion over the entire water column. Two main types of

breaker types have been studied: (1) the spilling breakers, where white foam, consisting of

a turbulent air/water mixture, appears at the wave crest and spills down the front face of the

propagating wave; and (2) the plunging breakers, where the front face of the steepening

wave overturns and impacts the forward face of the wave. These two breaker types have

been shown to have similar initial motions, but with different length scales [9]. When

approaching a beach, the waves change form due to the decrease in water depth. The

forward face of the wave steepens and the wave becomes asymmetric. Once the front face

becomes almost vertical, a jet of liquid is projected from the crest of the wave. The tongue

of water thrown from the crest develops and free falls down forward into a characteristic

overturning motion, and eventually hits the water at the plunge point. Depending on the

position of the plunge point, different breaker types can be observed. If the plunge point is

located very near to the crest of the wave, the resulting splash is directed down the wave

leading to a spilling breaker. Otherwise, if the jet is ejected farther towards the lower part

of the face of the steepening wave, the wave becomes a plunging breaker. The plunging jet

encloses an air pocket when it finally hits the wave face at the plunge point. The jet re-

enters the water after impact, forcing up a second jet, called splash-up. The early works of

Miller [132], Basco [9], Jansen [97] and Bonmarin [18] were dedicated to qualitative

description of the dynamics of the breaking process, the air entrainment and the evolution

of the large-scale geometric properties of bubble plumes. The overturning process, sub-

sequent overturning motion and plunging jet impact were described, resulting in the

identification and tracking of breaker vortices trajectories. Some information about the

evolution (size, shape and position) of the bubble plumes were also detailed. The jet-splash

54 Environ Fluid Mech (2017) 17:47–77

123

Page 9: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

cycles, occurring several times in a single breaker, have been shown to be responsible for

the generation of a sequence of large-scale vortices with a horizontal axis of rotation, some

of these eddies have been shown to be co-rotating vortices and some counter-rotating

vortices depending on the splash-up mechanism [18, 132]. Nadaoka et al. [142] detailed the

flow field under a turbulent bore propagating towards the shoreline, resulting from a

spilling breaking wave. Large-scale horizontal eddies are present in the bore front, while

behind the wave crest the flow structure changes rapidly into obliquely downward stretched

three-dimensional (3D) eddies, so-called ‘obliquely descending eddies’. Lin and Hwung

[120], Govender et al. [85] and Kimmoun and Branger [101] also described the large

motions of aerated regions under plunging breaking waves, with splash-ups and vortical

structures.

Miller [132] measured the average bubble concentration in plunging and spilling

breakers and indicated a larger bubble density presence in plunging breakers (about 31 %

in the late stage compared to 19 % for spilling breakers); these results were in agreement

with earlier descriptions from Miller [131]. Lamarre and Melville [109] concluded that a

large portion of the mechanical energy of the wave was lost in entraining the bubble

clouds. High values of void fractions (up to 100 %) were found next to the free-surface,

and void fractions of at least 20 % were observed for up to half a wave period after the

breaking occurrence. They later confirmed that air entrainment was closely related with the

energy dissipation of the breaking wave [111]. Several other works provided more com-

prehensive laboratory measurements of the void fraction in breaking waves, detailing the

vertical and horizontal distributions of void fraction. Cox and Shin [60] measured the void

fraction in the aerated region at a point using a capacitance probe, and observed peak

ensemble-averaged void fractions in the range of 15–20 %. Surprisingly, they measured

higher void fractions under the spilling breakers than under the plunging breakers. The

temporal variation of void fraction, above and below the still water level, was analysed

using three breaker types (spilling, spilling/plunging and plunging). The temporal variation

of void fraction above the still water normalized by the wave period and average void

fraction appears to be remarkably self-similar (independently of the breaker type). Hwung

et al. [94] found a deeper penetration of air bubbles under plunging breaking waves and

higher void fractions (18 %), compared to the spilling breaking waves (12 %). Similarly

Hoque and Aoki [90], using a conductivity probe, measured maximum void fractions of 20

and 16 % beneath plunging and spilling breakers, respectively. Mori et al. [135] obtained

void fractions of 19 % beneath spilling breaking waves and 24 % beneath plunging

breaking waves, using dual-tip resistivity void probe. Interestingly they also studied scale

effects according to Froude similarity and using two different scaled experiments. Void

fractions were affected by the geometric scale, with larger quantities being found in the

larger experiment, while the bubble size spectra proved to be nearly independent. Kim-

moun and Branger [101] estimated the evolution of void fractions using particle image

velocimetry images and velocity measurements. They reported large void fractions of up to

88 % in the first splash-up location, decreasing slowly when the breaking wave propagates

towards the shore, with values between 20 and 30 %. Much lower void fractions were

found in other studies. Kalvoda et al. [98] investigated the geometric and kinematic

characteristics of large air bubbles clouds produced by spilling breaking wind waves. They

observed that the lifetime of the bubble cloud was about 1.4 times the wave period, with

bubbles diameters in the range of 1.0–10 mm. They found a void fraction of about 0.4 %.

Leifer et al. [113] reported void fractions between 0.2 and 2.3 % beneath breaking wind

waves, using a video system to characterise the bubble clouds. Blenkinsopp and Chaplin

[14] studied plunging, spilling/plunging and spilling breaking waves. They calculated

Environ Fluid Mech (2017) 17:47–77 55

123

Page 10: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

integral properties of the bubble clouds and splash-ups, such as areas and volumes of air

entrained, trajectories of centroids and energy dissipation, and showed remarkable simi-

larity between plunging and spilling breakers. Their data indicated that the evolution of the

bubble clouds was subjected to scale effect. Rojas and Loewen [156] detailed the void

fraction evolution in spilling and breaking breakers. They observed that beneath plunging

breaking breakers, the mean void fraction ranged between 1.2 and 37 %, while beneath

spilling breaking waves, the mean void fraction ranged between 17 and 29 %. They found

that ‘‘an energetic spilling breaker may entrain approximately the same volume of air as a

steeper, larger-amplitude plunging breaker’’. They identified and tracked successive

bubble clouds, detailed the void fractions at each step of the breaking events, and found

that, beneath the spilling breaker, the celerity of the bubble cloud compared with the phase

speed. Beneath the plunging breaker, the celerity of the air cavity was about 70 % of the

phase speed. This has to be compared to the celerity of the bubble cloud entrained by the

propagating splash-up which has been measured to be about 90 % of the phase speed.

Blenkinsopp and Chaplin [14] and Rojas and Loewen [156] found that the volume of air

entrained by the splash-up, observed during a plunging breaking event, was greater than

the volume of air entrained by the initial plunging jet (about 60 % more). More recently

Anguelova and Huq [1] used an imaging technique to quantify the phase dependent void

fraction, and measured values reaching 80–99 % at the wave crest phase and decreasing to

20–30 % at the trough phase. Lim et al. [119] confirmed these results in the case of a

plunging breaker. They showed that the distribution of the turbulent intensities matched the

vorticity and void fraction fields. Nevertheless, some differences could be observed in the

experimental results for the peak values of void fraction, indicating a strong temporal and

spatial variability in the unsteady breaking waves [119]. The difference in the locations of

the measurements and the method used to generate the breaking could also be responsible

for the discrepancies. Some authors indicated that the mean void fraction could be mod-

elled by a linear function of time followed by an exponential decay. Hoque and Aoki [90]

found that the void fraction distribution followed the analytical solution of an advection

equation. This is not surprising as there is a general consensus about in void fractions

contours in the breaking waves, shapes and general kinematics of the aerated regions.

However, some differences can also be noted. Lamarre and Melville [109] found that that

the temporal variation of the normalized void fraction in deep water breakers could be

fitted by a power law t-2.3, while Rojas and Loewen [156] found the decay of void fraction

followed a power law t-2.6. Lim et al. [119] reported similar findings. Cox and Shin [60]

found a linear growth followed by exponential decay, but with a different rate, which could

be attributed to the shallow-water conditions used in their works.

3.2 Tidal bores

Undular tidal bores are observed for Froude numbers less than 1.3–1.4, and breaking tidal

bores with a marked roller are seen for Froude numbers large than 1.4–1.6 [37, 104].

Velocity measurements in breaking tidal bores were performed using particle image

velocimetry and acoustic Doppler velocimetry with most data obtained in the clear-water

column below the aerated roller region and for Froude numbers less than 2.5

[38, 69, 91, 105]. The data indicated a sharp deceleration of the flow associated with the

bore passage, as well as large fluctuations of all velocity components during and shortly

after the bore roller passage [105, 114].

Air–water flow measurements in tidal bores are rare. Leng and Chanson [114] high-

lighted some distinctive air bubble entrainment at the toe of the roller. Their data showed a

56 Environ Fluid Mech (2017) 17:47–77

123

Page 11: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

substantial number of bubbles with millimetre sizes: i.e., between 1 and 5 mm, with larger

bubbles detected at higher vertical elevations in a more intermittent manner. The same data

set showed that the toe perimeter shape fluctuated rapidly with transverse distance and

time, and large fluctuations in bore celerity were observed. That is, the celerity of the roller

toe fluctuated rapidly with both time and transverse distance, although in a quasi-two-

dimensional manner on average [115]. Yet both laboratory and field observations hinted to

the existence of large-scale three-dimensional bubbly structures [43, 114, 116].

3.3 Hydraulic jumps

A hydraulic jump is the sudden and rapid transition from a supercritical to subcritical flow,

characterised by the development of large-scale turbulence, surface waves and spray,

energy dissipation and air entrainment (Figs. 1, 2). Undular hydraulic jumps, observed for

Froude numbers less than 2–3, are not discussed herein. The breaking jump is a turbulent

shear flow [92, 150, 157], and vorticity is advected in the air–water shear zone. The first

successful air–water flow measurements were conducted by Rajaratnam [149]. Using hot

film anemometry, Resch and Leutheusser [153] and Babb and Aus [2] measured distri-

butions of Reynolds stresses and turbulence intensity in hydraulic jump rollers. High-

quality data were obtained using dual-tip phase-detection probes, allowing significant

improvements on the characterisation of the air–water flow field inside the jump roller.

Most studies were conducted for inflow Froude numbers greater than 5 [39, 49, 178].

Relatively Fewer studies investigated hydraulic jumps with lower Froude numbers

[29, 140]. Other research works also used flow visualization techniques [112, 137].

All experimental investigations in breaking jumps showed some intense air entrainment,

with singular air entrapment at the roller toe and interfacial aeration across the roller free-

surface [178]. Air is entrapped at the discontinuity between the impinging flow and the

roller: i.e., the roller toe (Figs. 1, 2). The roller toe perimeter is a line source of air bubbles,

as well as a line source of vorticity [31, 91]. In the shear region downstream of the roller

toe, the local maximum in void fraction Cmax decays longitudinally in a quasi-exponential

manner [49, 140]. A key feature is the developing shear layer and the recirculation region

above. The turbulent shear flow is somehow analogous to a wall jet although basic dif-

ferences between ideal wall jet and hydraulic jumps include the effect of longitudinal

pressure gradient qP/qx\ 0, conservation of mass, and the natural forcing which is likely

to enhance the two-dimensionality and periodicity of coherent structures [99]. In the jump,

the maximum velocity Vmax decays longitudinally as x-1/2 [39, 49, 139].

Figure 3 presents some typical air–water flow measurements in hydraulic jumps with

breaking roller. Figure 3a shows longitudinal variations in vertical distributions of void

fraction and interfacial velocity in the roller, where C is the time-averaged void fraction, V

is the interfacial velocity, x is the longitudinal distance from the roller toe, Lr is the roller

length, z is the vertical elevation, and d1 and V1 are the inflow depth and velocity

respectively. Overall the measurements highlighted two distinct air–water flow regions: (1)

the air–water shear layer, and (2) the upper free-surface region above. The shear layer is

characterised by strong interactions between the entrained air bubbles and turbulent

coherent structures, associated with a local maximum in void fraction, bubble count rate

and longitudinal velocity. In the shear layer, the distributions of void fractions follow an

analytical solution of the advection–diffusion equation for air bubbles assuming singular

air entrapment at the roller toe:

Environ Fluid Mech (2017) 17:47–77 57

123

Page 12: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

(A)

C/10, V/V1, (x-x1)/d1

z/d 1

-0.05 0.05 0.15 0.25 0.35 0.45 0.550

1

2

3

4CV/V1Z90/d1

(B)

C

z/d 1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

1

2

3

4

5

6

7(x-x1)/d1=15.2(x-x1)/d1=11.4(x-x1)/d1=7.59(x-x1)/d1=3.80

Singular aerationInterfacial aeration

(C)

V/Vmax

(z-Z

Vm

ax)/Z

0.5

-0.5 -0.3 -0.1 0.1 0.3 0.5 0.7 0.9 1.1 1.2-0.5

0

0.5

1

1.5

2

2.5

3Wall jetx/Lr=0.085x/Lr=0.169x/Lr=0.254x/Lr=0.394x/Lr=0.508x/Lr=0.677

58 Environ Fluid Mech (2017) 17:47–77

123

Page 13: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

C ¼qairq

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

4� p� D� x0p � exp �

z0�1ð Þ2x0

4� D

!

þ exp �z0þ1ð Þ2x0

4� D

! !

ð1Þ

where x0 = x/d1, z0 = z/d1, D is a dimensionless diffusion coefficient, q = V1 9 d1, and

qair is the entrained air flow rate per unit width. D and qair/q are derived from best data fit.

In the upper free-surface region above, the void fraction increases monotonically with

increasing distance from the bed from a local minimum up to unity, following an analytical

solution of the advection–diffusion equation for interfacial aeration/de-aeration:

C ¼ 1

2� 1� erf

1

2�

ffiffiffiffiffiffiffiffiffiffiffiffiffiffi

V1 � x

Dt

r

� z� Z50

x

� �� �

ð2Þ

where Dt is a diffusion coefficient, Z50 is the elevation where C = 0.5 and erf is the Error

function. Both Eqs. (1) and (2) are compared with experimental data in Fig. 3b, with solid

lines and dashed lines respectively. The interfacial velocity distribution presents a self-

similar profile close to that of wall jet:

V� Vrecirc

Vmax � Vrecirc

¼ exp �0:88� z� zVmax

z0:5

� �2 !

zVmax \z ð3Þ

where Vmax is the maximum velocity in the shear region, Vrecirc is the negative recircu-

lation velocity, z0.5 is the elevation where V = (Vmax ? Vrecirc)/2 and zVmax is the ele-

vation where V = Vmax. Equation (3) is compared with experimental data in Fig. 3c. Note

the negative time-averaged recirculation velocity, with Vrecirc/Vmax *-0.4 to -0.6 on

average [39, 185] (Fig. 3c). Further observations include the integral turbulent length scale

Lt characterising the size of large turbulent eddies advecting the bubbles in the hydraulic

jump roller [34, 179]. Typical results highlight the relationship between integral turbulent

length scale and inflow depth, with on average Lt/d1 * 0.5. Moreover, several studies

were dedicated to the analysis of the clustering process in a hydraulic jump [86], focussed

on the interaction between air bubbles and turbulence, and confirming the influence of the

bubble clustering on the surrounding flow field whereby hydrodynamic interactions and

enhanced velocity fluctuations can play an important role.

4 Discussion

Based on the previous section, we propose a discussion on limitations, disagreements and

obstacles which have to be overcome to improve the knowledge of the turbulent air–water

dynamics encountered in breaking waves and bores. We focus on the flow aeration and

bubble sizes distribution in the water column, water salinity and the definition of the

Froude number.

bFig. 3 Air bubble entrainment in a hydraulic jump with a marked roller. a Longitudinal variations ofvertical distributions of void fraction and interfacial velocity—Data set: Chachereau and Chanson [29],d1 = 0.044 m, Fr = 3.1, Re = 8.9 9 104. b Vertical distributions of void fraction—Data set: Chachereauand Chanson [29], d1 = 0.0395 m, Fr = 5.1, Re = 1.3 9 105. c Vertical distributions of interfacial velocityin the roller—Data set: Chanson [37], d1 = 0.018 m, Fr = 9.2, Re = 6.9 9 104

Environ Fluid Mech (2017) 17:47–77 59

123

Page 14: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

4.1 Flow aeration and bubble sizes

Very few detailed studies on air entrainment induced by breaking waves exist (Sect. 3.1).

A number of issues remain and definitive conclusions have not yet been drawn. Wave

conditions and characteristics are not easily comparable. For example, some studies were

centred on deep-water breaking waves while others focused on depth-limited conditions,

some breaking waves were mechanically generated, others were wind-waves. Further the

metrology (optical, acoustic, laser, conductivity BIV, PIV, video, photographs) and

experimental apparatus (single or double probes, measurements locations, assumptions for

analysis, mixture density estimations for wave characteristics, etc.) often suffered from

limitations and inaccuracies (e.g. [27, 33, 60]). Phase-detection needle probes are intrusive

point measurements, while BIV or PIV are non-intrusive but remain inaccurate when

operated in highly aerated turbulent flows (bubble clustering, shape changes, light

reflection, image processing technique and analysis). Sampling rate is another issue when

trying to characterise turbulent bubbly flow properties. A number of studies thus used

combined techniques to ensure a certain level of confidence in their measurements.

Several findings may be highlighted. All authors reported that the energy of the waves

was dissipated during the early stages of the breaking events, due to the splash-up kine-

matics, turbulence generation and air entrainment process. The kinetic energy was found to

decay as t-1, both experimentally [70, 109, 110, 130, 152] and numerically

[52, 95, 96, 108, 125, 126]. The same decay rate was observed for all breaker types

(spilling, weak plunging and plunging) [126, 130], while similar integral properties in

terms void fraction were comparable for different types of breaking waves [14, 60].

Locations of air entrapment in the roller can be identified at the toe of the advancing roller,

and on the whole interface of the roller, where multiple splashes are observed

[18, 119, 120, 132, 142, 156]. Large coherent eddies play a significant role in turbulent

production due to the large velocity gradients and contribute to shear in the mean flow and

fluctuations.

The highest velocities and turbulence intensities are found in the front part of the wave

crest, at the toe of the wave roller and in the shear regions. In both spilling and plunging

waves, the propagation of the roller to the shore line is associated with higher levels of

turbulence [85]. The breaking bore is widely recognized as being the primary source of

turbulence under breaking waves. The void fraction appeared to be correlated with the flow

turbulence and turbulent kinetic energy [60, 135], but experimental results have not yet

proved that the turbulent properties are enhanced by high concentrations of bubbles, and

vice and versa. Indeed, the relations between void fraction and bubble size distributions,

and between void fraction and turbulence, remain unclear: ‘‘the entrained bulk of air and

bubbles generate turbulence which breaks coarser bubbles into finer ones in this phase’’

[136]. Gemmrich [82] speculated that the bubble fragmentation process is mainly driven by

turbulence. But the bubble dynamics might also be a major source of turbulence in the

surrounding fluid.

Altogether breaking waves can produce large millimetre bubbles, although smaller

bubbles may be found in the breaking region (Table 1). Table 1 summarises a few detailed

observations. Larger bubbles are mainly present in the roller region, while much smaller

bubbles with large residence times may be found deeper. Bubble populations under

breaking waves are very inhomogeneous in time and space. Bubbles can be observed to

split under several breakup mechanisms, including turbulent induced breakup, shear-driven

breakup resonant oscillation and tip-streaming [35, 57, 169]. The dynamics of a bubble

60 Environ Fluid Mech (2017) 17:47–77

123

Page 15: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

under unsteady perturbations is affected by the balance between forces acting on the

bubble, but the residence time of the bubble in the turbulent flow field should also be taken

into account [31, 81]. A large amount of bubbles are generated and broken in the shear

zones, where velocity gradients play a great role in the breakup process. Current breakup

models are unable to quantify accurately the breakup rate and bubble size distributions in a

complete unsteady turbulent flow with high void fractions, partly because the two-phase

flow mixture beneath breaking waves is a highly unsteady process, never reaching an

equilibrium. As an illustration, Fig. 4 presents the normalized probability distribution

functions of bubble radii beneath a plunging seawater jet, at three vertical elevations. In

Fig. 4, each histogram column represents the probability of bubble radius in 0.25 mm

increments: that is, the probability of a bubble radius from 1.0 to 1.25 mm is represented

by the column labelled 1.0. The last column (i.e. 0.5) indicates the probability of bubble

radii large than 5 mm. The data showed a broad spectrum of bubble sizes at all elevations,

from less than 0.25 mm more than 5 mm (Fig. 4). The bubble radius distributions showed

a preponderance of small bubble sizes relative to the mean: the mode was between 0 and

0.5 mm, although the mean radius was about 1.5–2.3 mm (Table 1). If equilibrium exists,

the ‘equilibrium’ bubble size distribution is dictated by all the events taking place in the

flow, in particular by the relative rates of bubble breakup and bubble coalescence, the

bubble-turbulence interactions and the bubble clustering.

Table 1 Summary of bubble size observations

Flowtopology

Description Equivalent bubbleradius (mm)

Reference

Breakingwaves

Open ocean (Fig. 14) 0.8–1 Bowyer [20]

Open ocean (Fig. 6)Laboratory focused waves paddle (10 cmplunging) (Fig. 4)

1 Dean and Stokes[67]

Open ocean 0.7 Stokes et al. [166]

LaboratoryWind ? paddle (*12 cm breaking height)Fresh (but not clean) water

1–1.7 Leifer et al. [113]

Laboratory focused waves paddle (*20 cmplunging)

Fresh water

1.5 Rojas and Loewen[156]

Laboratory wave paddle sloping beachspilling—plunging (12.6–16.5 cm mid-scale/5.0–6.5 small-scale)

2.01 (mid-scale)–4.28(small-scale)

Mori et al. [135]

Laboratory wave paddle submerged reefspilling—plunging (9.5 cm)

Tap water

1.7–2.4 Blenkinsopp andChaplin [15]

Plungingjet

Circular plunging jet (Ø = 25, 12.5 and6.85 mm)

Tap water (Table 3)

4.7 (Ø = 25 mm)2.2–3.7(Ø = 12.5 mm)

1.7–2 (Ø = 6.85 mm)

Chanson et al. [32]

Circular plunging jet (Ø = 12.5 mm)Tap water, saltwater, seawater (Table 5)

2.2–3.1 (tap water)2.3–3.3 (saltwater)1.5–2.3 (seawater)

Chanson et al. [48]

Environ Fluid Mech (2017) 17:47–77 61

123

Page 16: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

4.2 Water solution: freshwater, saltwater or seawater?

A number of studies tested the influence of salinity on the air bubble entrainment, although

there is no agreement between most published results. A majority of such studies tested

experimentally saltwater solutions, typically by gradually adding salt to freshwater: i.e.,

synthetic seawater [20, 28, 122, 160] (Also Table 1). Differences have been attributed to

some form of inhibited coalescence process in saltwater. Controlled experiments with

genuine seawater are rare (e.g. [16, 25, 48]). In strongly aerated regions (void fractions

[2 %), the results showed differences between freshwater, saltwater (synthetic seawater)

and genuine seawater in terms of void fraction and bubble sizes. Despite differences in

trends, an overall conclusion was that large amount of air bubbles were entrained in all

solutions, and the majority of bubbles in the aerated flow region had radii on the order of a

millimetre (Fig. 4).

It must be stressed however that the rare data were obtained with different geometric

scales. In air–water flows, experimental results demonstrated that scale effects may be

significant [31, 42, 45, 151, 184], in particular in terms of bubble numbers and sizes. In

turn, small scale experiments are more sensitive to changes in water salinity, and this may

bias the findings.

4.3 Froude number and definition

The definition of the Froude number is of paramount importance, because it is used to

detect the breaking event onset and termination during the evolution of the waves in the

surf zone. It is also the critical parameter allowing scaling of the data, assuming analogies

between hydraulic jumps, tidal bores and rollers due to breaking waves. Table 2 regroups a

number of definitions. Several criteria have been used in the past to detect the inception of

wave breaking. Wave steepness, based on Stokes’ limiting wave height, has been widely

used as an proxy for wave breaking. The occurrence of breaking has also been related to

the energy content in the waves, to the acceleration of the free-surface, or to the

Bubble radius (mm)

FDP

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 50

0.03

0.06

0.09

0.12

0.15

0.1810 mm - 3673 bubbles15 mm - 3080 bubbles25 mm - 1808 bubbles

Fig. 4 Pseudo-bubble radius distributions in seawater generated by a vertical circular plunging jet, at 10, 15and 25 mm beneath the free-surface—Data set: Chanson et al. [48], jet diameter: 12.5 mm, impact velocity:3.1 m/s, water solution: Pacific Ocean seawater off Japan coastline—All bubble sizes were recorded in air–water regions with void fractions greater than 2 %

62 Environ Fluid Mech (2017) 17:47–77

123

Page 17: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

prescription of a limiting value to the fluid velocity at the crest of the wave prior to

breaking (horizontal fluid velocity: [146] (see Fig. 2); [181]; vertical fluid velocity: [103]).

However consensus and agreement has not been reached for the establishment of a simple,

accurate and universal valid criterion. This is in part because wave breaking is a highly

complicated process involving highly non-linear effects including varying bathymetries,

currents, and wind (not taken into account in any study). An improvement in the quest has

been the use of a Relative Trough Froude Number (RTFN, [145]), based on the analogy

with a moving hydraulic jump in a one-dimensional open channel flow. The Froude

number can be obtained, as detailed in Table 2, by the mean of the P-theorem, momentum

considerations (Appendix 2) and by simple construction considering the quantities avail-

able. A discussion is provided in Appendix 2, since consistency and physical adequacy are

key features of this study. It is also important to consider the quantities available in the

various sources of data (Table 2, third column) (Fig. 5).

The strength of a breaking roller and the rate of energy dissipation in the roller are

directly linked to the Froude number [5, 21]. This is detailed in Appendix 2. The roller

length Lr increases monotonically with increasing Froude number, as illustrated in Fig. 6a.

For stationary hydraulic jumps, very detailed experimental measurements yielded:

Table 2 Froude number definitions in breaking waves, bores, surges and jumps

Froude numberdefinition

Reference Quantities Comment

Fr = (U ? V1)/(g 9 d1)

1/2Bazin [11],Boussinesq[19]

Implies the knowledge of thecurrent condition V1 and d1flowing adversely the roller,together with the celerity of theroller toe U

Classical definition, satisfying themass and momentumconservation equations, in arectangular horizontal channel

Fr = (U ? V1)/(g 9 A1/B1)

1/

2

Chanson [41] Implies further the knowledge ofthe initial cross-section area A1

and free-surface width B1

Application of mass andmomentum conservationequations for an irregular-shapedchannel

Fr = (ccrest -utrough)/ctrough

Okamoto andBasco [145]

Velocity of the crest, trough anddepth averaged velocity in thetrough region

Can be applied to deep waterbreaking waves

Fr = (((2 9 d2/d1 ? 1)2 -1)/8)1/2

Tissier et al.[171]

Water heights upstream anddownstream the jump

Not accurate enough [3]

Fr = (1 ? 3/2 9 Hr/d1 ? 1/2 9 (Hr/d1)

2)1/2

Tonelli andPetti [172]

Same as the previous definition,but written as a function of theroller height and water depth infront of the roller

Not accurate enough [3]

Fr = us/cb Bacigaluppiet al. [3]

Cheng andWang [53]

us is the free surface velocity at thecrest of the wave, and cb thewave celerity

Implies calibration. Fr = 0.75 to 1found to be the value whenwaves start breaking [3, 53].

Problem: based on the idea thatwave breaking occurs when thefree surface velocity exceeds thewave celerity (which is not truefor spilling, and true for thecelerity inside the crest forplunging)

Environ Fluid Mech (2017) 17:47–77 63

123

Page 18: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

Lr

d1¼ 6� Fr� 1ð Þ ð4Þ

where d1 is the upstream depth and Fr = (U ? V1)/(g 9 d1)1/2 (see symbol definition in

Fig. 1). Equation (4) is compared to stationary hydraulic jump and tidal bore data in

Fig. 6a. Practically another relevant quantity is the length of the aerated region Lair.

Typical experimental observations are reported in Fig. 6b, suggesting that the aeration

region length may be about 1.5–2 times longer than the roller itself.

Another measure for the roller strength is the height of the roller hr. Indeed the height of

the roller increases with increasing Froude number. Basic momentum considerations show

a unique relationship between roller height and Froude number for Fr[ 2 (Appendix 2):

hr

d1¼ 1

2�

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

1þ 8� Fr2p

� 3� �

Fr [ 2 ð5Þ

For Fr\ 2, physical observations in tidal bores showed that the surface immediately

upstream of the roller was curved upwards and the roller height was less than the difference

in conjugate depths (Fig. 7). This is illustrated in Fig. 7, by comparing experimental data

and Eq. (5). For breaking bore (i.e. 1.4\ Fr\ 2), the roller height data were best cor-

related by

hr

d1¼ 1:59� Fr� 1:16ð Þ0:94 1:4 \Fr \2 ð6Þ

The mean slope of the breaking roller can be obtained tan/ = hr/Lr. When the height of

the breaker hr is lower than the difference in conjugate depths (d2 - d1), the front face of

the breaking wave may not be fully saturated. Importantly the evolution of the roller is

highly unsteady and depends on the evolution of many parameters, including the variation

of the bathymetry or the current for example. Hence some simplistic parameters such as the

roller height might not characterise accurately the breaking wave roller.

Lr/hr

snoitavresb ofob

N

0 1 2 3 4 5 6 7 8 90

2

4

6

8

10

12

14

16Steady breakerTidal boreHydraulic jump

Fig. 5 Probability distributionsfunctions of roller aspect ratio Lr/hr—Re-analysis of large-scalewave breaking experiments [58,87], stationary hydraulic jumpdata [106, 138, 139, 178, 180],and tidal bore data [51, 138]

64 Environ Fluid Mech (2017) 17:47–77

123

Page 19: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

Fr

L r/d

1

1 2 3 4 5 6 7 8 9 10 11 120

10

20

30

40

50

60

70

80

90

100

110

Tidal bore laboratoryTidal bore SéluneMurzyn et al (2007)Kucukali and Chanson (2008)Murzyn & Chanson (2009)Richard & Gavrilyuk (2013)Wang et al. (2015)Wang and Chanson (2015)6×(Fr-1)

Chanson (2010)Virey (2013)2.91×(d2/d1-1)

(B)

Fr

L air/

d 1

1 2 3 4 5 6 7 8 9 10 11 120

10

20

30

40

50

60

70

80

90

100

110

6×(Fr-1)Chanson (2010)Chachereau and Chanson (2011)Virey (2013)

(A)

Fig. 6 Roller length and aeration length in breaking hydraulic jumps and tidal bores as functions of theFroude number (Fr = (V1 ? U)/(g 9 d1)

1/2. a Roller length in breaking hydraulic jumps and tidal bores—Most data are hydraulic jump data unless indicated in the legend; tidal bore data by Chanson and Toi [51].b Aeration region length in hydraulic jumps

Environ Fluid Mech (2017) 17:47–77 65

123

Page 20: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

5 How can better flow analogies help?

This study clearly shows that more work needs to be done to elucidate the physics of the

unsteady motion of a breaking roller. The metrology is clearly an issue to get accurate and

reliable physical measurements in this rapidly-varied aerated turbulent flow. The scale

effects are also a point to be addressed, with a close look at the differences due to the water

salinity (tap water, saltwater, seawater). Some practice, such as the use of the Hinze’s [89]

scale, are also believed to be misleading, as the theory was originally developed under

assumptions not valid in the case of breaking waves. The analogy between hydraulic

jumps, breaking tidal bores and breaking waves may be more appropriate in order to

deduce quantities aimed at a modelling effort.

A stationary hydraulic jump is the sudden transition from a high-velocity free-surface flow

impinging into a slower open channel flow. The rapid transition, called the roller, is char-

acterised by spray and splashing with a highly fluctuating free-surface, together with highly-

aerated turbulent flow structures within, and a large amount of energy dissipation takes place

[5]. The application of the energy principle across the roller gives an expression of the energy

dissipation. For a smooth horizontal rectangular channel and assuming hydrostatic pressure

upstream and downstream of the roller, the rate of energy dissipation is [4, 21]:

DEE1

¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

1þ 8� Fr2p

� 3� �3

16�ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

1þ 8� Fr2p

� 1� �

� 1þ 12� Fr2

� �

ð7Þ

where E1 is the upstream total energy: E1 = d1 ? (V1)2/(2 9 g). The rate of energy dis-

sipation may exceed 70 % for Fr[ 9 [32, 88]. However a non-breaking undular jump is

observed at low Froude numbers. The transition from a stationary undular jump into a jump

with a marked breaking roller takes place with the progressive apparition of characteristic

three-dimensional flow features with increasing Froude numbers, including cross-waves

Fr

h r/d

1, (d

2-d 1

)/d1

1.2 1.4 1.6 1.8 2 2.20

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

hr/d1 Leng & Chanson (2015)hr/d1 Koch & Chanson (2009)hr/d1 Chanson (2010)hr/d1 Docherty & Chanson (2012)(d2-d1)/d1=0.5×(sqrt(1+8×Fr2)-3)1.587×(Fr{-1.1612)0.9426

Fig. 7 Roller height in breakingtidal bores as function of theFroude number (Fr = (V1 ? U)/(g 9 d1)

1/2—Comparisonbetween breaking tidal bore dataand Eq. (5) (black solid line) andEq. (6) (red dashed line)

66 Environ Fluid Mech (2017) 17:47–77

123

Page 21: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

and cockscomb roller [46, 144]. A hydraulic jump with a fully-developed breaking roller is

observed only for Froude numbers greater than 2–4 [35, 158]. Thus a stationary jump

might not be truly representative of a spilling breaker, typically observed for Fr\ 2.

It is believed that a more relevant analogy is the tidal breaking bore, as seen in Fig. 2b.

A breaking bore with a quasi-two-dimensional roller is observed for Froude numbers

greater than 1.4–1.6 [104, 115], although localised form of breaking might appear for

Froude numbers above 1.3–1.4, including shock waves upstream of and limited breaking at

the first wave crest [37, 173]. In the field, tidal bores are typically observed for Froude

numbers less than 2.5, as shown in Table 3. As a further illustration, the Froude number

was estimated to be about 2 for the bore seen in Fig. 2b2. Based upon basic energy

considerations, the power dissipated in the bore roller may be estimated as (Appendix 2):

Pd ¼ q� g3=2 � d5=21 � L� Fr �

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

1þ 8� Fr2p

� 3� �3

16�ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

1þ 8� Fr2p

� 1� � ð8Þ

where Pd is in Watts, L is the transverse length of the bore roller and where

E1 = d1 ? (V1 ? U)2/(2 9 g) for a tidal bore. For the tidal bore seen in Fig. 2b2, the power

dissipated was about: Pd &160 MW. When a tidal bore propagates upstream in a sloping

channel, experimental observations showed a complex transformation of a bore into a sta-

tionary hydraulic jump [30, 40]. In laboratory, the transformation could last from 5 to 10 min

up tomore than 20 min. At full scale, the duration of the processwould bemuch longer, based

upon a Froude similitude scaling, but the physical process remains poorly understood.

Even if the external forces generating the final roller are different in the flows considered in

this study, the general aspect of the turbulent aerated bores seem visually similar in terms of

lengths scales, intensities and aeration. Indeed, looking at the general hydrodynamic features of

the flows presented in the literature, the roller region results in a sequence of splash-ups,

inducing large-scale horizontal vortices. Considering the description of the splashing hydro-

dynamics in breaking waves (e.g. [18, 119, 132]), the roller region is roughly composed of 3–4

large horizontal vortices, which is consistent with experimental observations in tidal bores for

Fr\ 2. All the experimental studies, whatever the flow, indicated the capacity of successive

Table 3 Detailed field observations of tidal bores

Reference River/Estuary Date Bore type d1 (m) V1

(m/s)U(m/s)

Fr

Wolanski et al. [183] Daly 2/07/2003 Undular 1.50 0.15 4.7 1.04

Simpson et al. [163] Dee 6/09/2003 Breaking 0.72 0.15 4.1 1.79

Chanson et al. [50] Garonne 10/09/2010 Undular 1.77 0.33 4.5 1.30

11/09/2010 Undular 1.81 0.30 4.2 1.20

Mouaze et al. [138] Selune 24/09/2010 Breaking 0.38 0.86 2.0 2.35

25/09/2010 Breaking 0.33 0.59 2.0 2.48

Reungoat et al. [154] Garonne 7/06/2012 am Undular 2.72 0.68 3.8 1.02

7/06/2012 pm Undular 2.65 0.59 4.6 1.19

Furgerot et al. [80] See 7/05/2012 Undular 0.90 0.40 3.2 1.39

Reungoat et al. [155] Garonne 13/10/2013 Undular 2.05 0.26 4.3 1.27

Leng and Chanson [115] Qiantang 6/09/2013 Breaking 1 – 3.6 2.1

Chanson [43] Qiantang 12/10/2014 Breaking 2 to 2.5 – 4.5 2.00

Italic data rough estimate; (–) data not available

Environ Fluid Mech (2017) 17:47–77 67

123

Page 22: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

splash-ups to entrap air large quantities of air. The co-existence between the large-scale

clockwise vortices and some smaller counter-clockwise vortices, generated by backward

splashing and impingement occurring in between co-rotative vortices pairs, is reported in all the

observations. These large eddies are responsible for the turbulent fluctuations of the shapeof the

rollers.All the observations indicated that both air entrainment, turbulent shear and vorticity are

generated along the surface of the rollers and in the toe region. The total air volume is less

important than the number of bubbles and their actual size distributions in the roller region. The

study of the highly fluctuating dynamics of the roller region should then bring fruitful insight in

the dynamics of the flow. The temporal evolution of this region is highly required to get a better

description of the variation of the roller height and energy dissipation, as a function of time.

The quantification of some parameters may be improved and some physical description can

be highlighted, using the most appropriate analogy (roller/tidal bore) rather than the one typ-

ically used (roller/hydraulic jump). The roller length and height have been found to be fairly

similar between hydraulic jumps and tidal bores (Fig. 6), even comparing laboratory and field

studies, thus free of scale effects. Considering the tidal bore analogy and the corresponding

observations presented in Fig. 5, the length of the roller should be considered within 1.5\Lr/

hr\ 5.The transitionbetweenbreakingandnon-breakingwaveswould takeplace for Fr[ 1.3,

as observed in tidal bores. It was shown that the same limit value was inaccurately attributed to

the transition between undular and breaking hydraulic jumps, but most recent studies proved it

to be for Fr[ 2 in stationary jumps. It is also interesting to notice that the celerity of the bore

region considered in tidal bores is the celerity of the toe [114, 115], and not the celerity of the

crest as used in modelling studies. The celerity of a tidal bore is found to be different from the

celerity of a solitary wave. In a wave current problem, the Froude number represents the ratio

between the velocity of a wave travelling and that of the current. The definition of the Froude

number, as given for tidal bores in Table 2, should then be useful for breakingwaves studies. It

is similar to the estimation based on the RTFN concept [62, 145], and should be appropriate

since itwould be suitable for any type of analogy (deep and shallowwater breakingwaves). It is

noted that the celerity of the crest, considered byOkamoto and Basco [145] should be replaced

by the celerity of the toe, ctrough being V1 with the current notation.

This study also presents some acknowledged limitations. The current work cannot provide

any definition of a criterion for breaking onset. Experimental studies do not focus on the

initiation and evolution of the breaking region in tidal bores or hydraulic jumps. Estimations of

the wave velocity during breaking can lead to substantial discrepancies [146] and remains a

challenge. A steepening wave can propagate for a long distance before clearly breaking and

developing a roller; so the crest kinematics still need to be elucidated in complex situations

(3D, current, turbulence, wind, etc.). The transient evolution of the roller is crucial for future

studies [115]. Future interests should also encompass bubble size distribution and dynamics,

two-phase gas–liquid turbulence and vorticity generation in the roller region. Interestingly,

tidal bore is itself also a great example of highly non-linear phenomenon, as the free-surface

evolution of the undulations observed in rivers and laboratories do not show any accordance

with classical wave theories (cnoidal, solitary, Stokes). Analogies are complicated both ways.

6 Conclusions

In this study, we analysed and discussed the different analogies proposed in the literature tomodel

the breaking process of the waves in the surf zone. Numerical models not only need experimental

data for rigorousandextensivevalidations, but also fordevelopment.Even themost up-to-date and

68 Environ Fluid Mech (2017) 17:47–77

123

Page 23: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

accurate numerical models are still limited by empirical aspects. So a better knowledge of the

temporal and spatial evolution of the aerated region under breaking waves is crucial.

Thequantitieswhichareneeded formost numericalmodels are acharacteristicFroudenumber

to detect the breaking process occurrence, the geometrical and dynamical characteristics of the

roller (height, length, mean front slope angle, celerity), the energy dissipation and the bubble size

distribution. Two features are reported to be of great importance in all experimental studies: the

bubble size distribution and the bubble cloud void fraction, the latter being highly dependent on

the accurate quantification of the number of bubbles. But this aspect of the flow remains a

technical challenge for both the numericalmodels and experimental technics. A number of issues

remain and definitive conclusions cannot be drawn. The actual best practices have been discussed

and better estimations have been proposed based on recent experimental data sets.We showed in

particular that measurements, performed in tidal bores will lead to improve the description of the

bubble plumes. The authors believe that the unsteady evolution of a decelerating tidal bore,

showing transition between a tidal bore to a stationary hydraulic jump [40], could bring intriguing

insights into the processes involved in the highly non-linear problem of breaking waves.

Acknowledgments The authors thank their students, former students and co-workers for their work andinput. The financial support of the Australian Research Council (Grant DP120100481) is acknowledged.

Appendix 1

See Table 4.

Table 4 Breaking waves, bores, surges and jumps: flow analogies in the literature

Analogy spilling breaker and steady breakers Banner and Phillips [6, 7], Duncan [73, 74],Banner and Peregrine [8]Cointe and Tulin [59]Lin and Rockwell [121]Dabiri and Gharib [63]

Analogy breaking wave and Plunging jet Cipriano and Blanchard [56]Hubbard et al. [93]Chanson and Cummings [44]Oguz et al. [143]Chanson et al. [47, 48]Salter et al. [160]

Analogy spilling breaker and hydraulic jump Longuet-Higgins [123]Peregrine and Svendsen [147]Madsen [128]Brocchini and Peregrine [23, 24]

Analogy spilling breaker and bore/positivesurge

Longuet-Higgins [123]Peregrine and Svendsen [147]Brocchini and Peregrine [24]

Analogy breaking bore/positive surge andhydraulic jumps

ALL Hydraulics Textbooks, dealing with open channelflows, as well as:

Darcy and Bazin [65]Stoker [164]Tricker [174]Henderson [88]Lighthill [118]Chanson [32, 41]

Other analogies—spilling breaker and densitycurrent

Longuet-Higgins and Turner [124]

Environ Fluid Mech (2017) 17:47–77 69

123

Page 24: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

Appendix 2: Basic momentum and energy considerations for boresand jumps

A bore is an abrupt rise in water depth (Fig. 1) and its front may be analysed as a hydraulic

jump in translation [41, 118, 174]. In a system of reference in translation with the bore

front, the integral form of the continuity and momentum equations gives a relationship

between the cross-section area upstream and downstream of the roller, A1 and A2

respectively, and the upstream Froude number Fr1 [114]

Fr21 ¼1

2� A2

A1

� B1

B� 2� B0

B

� �

þ B0

B� A2

A1

� �

þ A2

A2 � A1

� Ffric �W � sinh

q� g� A21

B

ð9Þ

where Fr1 is the bore Froude number: Fr = (U ? V1)/(g 9 A1/B1)1/2, V is the cross-

sectional averaged velocity positive downstream as shown in Fig. 1, U is the bore celerity

positive upstream, the channel cross-sectional area A measured perpendicular to the flow

direction, q is the water density, g is the gravity acceleration, Ffric is the friction force, W is

the weight force, h the angle between the bed slope and horizontal, and the subscripts 1 and2 refer to the flow conditions immediately before and after the bore roller respectively

(Fig. 1). In Eq. (9), B and B0 are characteristic widths functions of the bathymetry:

B ¼ A2 � A1

d2 � d1ð10Þ

B0 ¼

R

A2

A1

R

q� g� d2 � zð Þ � dA

12� q� g� d2 � d1ð Þ2

ð11Þ

with d the flow depth (Fig. 1b). For a smooth rectangular horizontal channel, Eq. (9) yields

to the Belanger equation:

d2

d1¼ 1

2�

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

1þ 8� Fr21

q

� 1

� �

Smooth horizontal rectangular channel ð12Þ

where d1 and d2 are respectively the upstream and downstream flow depth, and Fr1 sim-

plifies into: Fr1 = (U ? V1)/(g 9 d1)1/2 (Fig. 1). First presented in 1841 [12, 36], Eq. (12)

is inappropriate in an irregular channel [41]. The roller height hr is in first approximation:

hr = d2 - d1 and Eq. (12) may rewritten as:

Fr21 ¼1

2� d2 � d1

d1

� �2

þ 3

2

d2 � d1

d1Smooth horizontal rectangular channel ð13Þ

Note that the above development (Eq. 9, 12 and 13) assumes implicitly that the bore

celerity is uniform and the roller shape two-dimensional. Field and laboratory observations

suggested that these assumptions are simplistic [114].

The application of the energy principle across the bore roller gives an expression of the

energy dissipation [4, 21]:

70 Environ Fluid Mech (2017) 17:47–77

123

Page 25: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

DE ¼ d1 þV1 þ Uð Þ2

2� g� d2 +

V2 þ Uð Þ2

2� g

!

ð14Þ

assuming hydrostatic pressure upstream and downstream. For a smooth horizontal rect-

angular channel, the rate of energy dissipation becomes [5]:

DEE1

¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

1þ 8� Fr21

q

� 3

� �3

16�ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

1þ 8� Fr21

q

� 1

� �

� 1þ 12� Fr21

� �

ð15Þ

where E1 = d1 ? (V1 ? U)2/(2 9 g). The rate of energy dissipation ranges from 0 for

Fr1 = 1 to more than 70 % for Fr1[ 9 [32, 88]. The power dissipated in the bore is:

Pd ¼ q� g� V1 þ Uð Þ � d1 � L� DE ð16Þ

where L is the transverse length of the bore roller.In the above equations, the solution for a

stationary hydraulic jump is obtained for U = 0.

References

1. Anguelova MD, Huq P (2012) Characteristics of bubble clouds at various wind speeds. J Geophys Res117:C03036. doi:10.1029/2011JC007442

2. Babb AF, Aus HC (1981) Measurements of air in flowing water. J Hydraul Div ASCE107(12):1615–1630

3. Bacigaluppi P, Ricchiuto M, Bonneton P (2014) Upwind stabilized finite element modelling of non-hydrostatic wave breaking and run-up. Research Report INRIA, RR-8536, https://hal.inria.fr/hal-00990002v3

4. Bakhmeteff BA (1912) O Neravnomernom Dwijenii Jidkosti v Otkrytom Rusle. (Varied Flow in OpenChannel.) St Petersburg, Russia (in Russian)

5. Bakhmeteff BA (1932) Hydraulics of open channels, 1st edn. McGraw-Hill, New York6. Banner ML, Phillips OM (1974) On the incipient breaking of small scale waves. J Fluid Mech

65:647–6567. Banner ML, Melville WK (1976) On the separation of air flow over water waves. J Fluid Mech

77:825–8918. Banner ML, Peregrine DH (1993) Wave breaking in deep water. Annu Rev Fluid Mech 25:373–3979. Basco DR (1985) A qualitative description of wave breaking. J Waterw Port Coast Ocean Eng

111(2):171–18810. Battjes JA (1988) Surf-zone dynamics. Ann Rev Fluid Mech 20:257–29311. Bazin H (1865) Recherches Experimentales sur la Propagation des Ondes. (‘Experimental Research on

Wave Propagation.’) Memoires presentes par divers savants a l’Academie des Sciences, Paris, France,Vol. 19, pp. 495–644 (in French)

12. Belanger JB (1841) ‘‘Notes sur l’Hydraulique.’’ (‘Notes on Hydraulic Engineering.’) Ecole Royale desPonts et Chaussees, Paris, France, session 1841–1842, 223 pp (in French)

13. Bjørkavag M, Kalisch H (2011) Wave breaking in Boussinesq models for undular bores. Phys Lett A375:1570–1578

14. Blenkinsopp CE, Chaplin JR (2007) Void fraction measurements in breaking waves. Proc R Soc A463:3151–3170. doi:10.1098/rspa.2007.1901

15. Blenkinsopp CE, Chaplin JR (2010) Bubble size measurements in breaking waves using optical fiberphase detection probes. J Ocean Eng 35(2):388–401

16. Blenkinsopp CE, Chaplin JR (2011) Void fraction measurements and scale effects in breaking wavesin freshwater and seawater. Coast Eng 58:417–428

17. Bombardelli FA (2012) Computational multi-phase fluid dynamics to address flows past hydraulicstructures. Proceedings of 4th IAHR International Symposium on Hydraulic Structures, APRH—

Environ Fluid Mech (2017) 17:47–77 71

123

Page 26: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

Associacao Portuguesa dos Recursos Hıdricos (Portuguese Water Resources Association), J. Matos, S.Pagliara & I. Meireles Eds., 9–11 February 2012, Porto, Portugal, Paper 2, 19 pages (CD-ROM)

18. Bonmarin P (1989) Geometric properties of deep-water breaking waves. J Fluid Mech 209:405–43319. Boussinesq JV (1877) Essai sur la Theorie des Eaux Courantes. (‘Essay on the Theory of Water

Flow.’) Memoires presentes par divers savants a l’Academie des Sciences, Paris, France, Vol. 23,Serie 3, No. 1, supplement 24, pp. 1–680 (in French)

20. Bowyer PA (2001) Video measurements of near-surface bubble spectra. J Geophys Res106(C7):14179–14190

21. Bresse JA (1868) Cours de Mecanique Appliquee Professe a l’Ecole Imperiale des Ponts et Chaussees.(Course in Applied Mechanics lectured at the Pont-et-Chaussees Engineering School.) Gauthier-Villars, Paris, France, 586 p (in French)

22. Brocchini M (2013) A reasoned overview on Boussinesq-type models: the interplay between physics,mathematics and numerics. Proc R Soc A 469:20130496

23. Brocchini M, Peregrine DH (2001) The dynamics of strong turbulence at free surfaces. Part 1.Description. J Fluid Mech 449:225–254

24. Brocchini M, Peregrine DH (2001) The dynamics of strong turbulence at free surfaces. Part 2. Free-surface boundary conditions. J Fluid Mech 449:255–290

25. Callaghan AH, Stokes MD, Deane GB (2014) The effect of water temperature on air entrainment,bubble plumes, and surface foam in a laboratory breaking-wave analog. J Geophys Res Ocean. doi:10.1002/2014JC010351

26. Carrica PM, Drew D, Bonetto F, Lahey RT Jr (1999) A polydisperse model for bubbly two-phase flowaround a surface ship. Int J Multiph Flow 25:257–305

27. Cartellier A, Achard JL (1991) Local phase detection probes in fluid/fluid two-phase flows. Rev SciInstrum 62(2):279–303

28. Cartmill J, Su M (1993) Bubble size distribution under saltwater and freshwater breaking waves. DynAtmos Oceans 20:25–31. doi:10.1016/0377-0265(93)90046-A

29. Chachereau Y, Chanson H (2011) Bubbly flow measurements in hydraulic jumps with small inflowfroude numbers. Int J Multiph Flow 37(6):555–564. doi:10.1016/j.ijmultiphaseflow.2011.03.012

30. Chanson H (1996) Free-surface flows with near-critical flow conditions. Can J Civ Eng23(6):1272–1284

31. Chanson H (1997) Air bubble entrainment in free-surface turbulent shear flows. Academic Press,London

32. Chanson H (2004) The hydraulics of open channel flow: an introduction, 2nd edn. Butterworth-Heinemann, Oxford

33. Chanson H (2005) Air–water and momentum exchanges in unsteady surging waters: an experimentalstudy. Exp Thermal Fluid Sci 30(1):37–47

34. Chanson H (2007) Bubbly flow structure in hydraulic jump. Eur J Mech B/Fluids 26(3):367–384.doi:10.1016/j.euromechflu.2006.08.001

35. Chanson H (2009) Turbulent air–water flows in hydraulic structures: dynamic similarity and scaleeffects. Environ Fluid Mech 9(2):125–142. doi:10.1007/s10652-008-9078-3

36. Chanson H (2009) Development of the Belanger equation and backwater equation by Jean-BaptisteBelanger (1828). J Hydraul Eng ASCE 135(3):159–163. doi:10.1061/(ASCE)0733-9429(2009)135:3(159)

37. Chanson H (2010) Undular tidal bores: basic theory and free-surface characteristics. J Hydraul EngASCE 136(11):940–944. doi:10.1061/(ASCE)HY.1943-7900.0000264

38. Chanson H (2010) Unsteady turbulence in tidal bores: effects of bed roughness. J Waterw Port CoastOcean Eng ASCE 136(5):247–256. doi:10.1061/(ASCE)WW.1943-5460.0000048

39. Chanson H (2010) Convective transport of air bubbles in strong hydraulic jumps. Int J Multiph Flow36(10):798–814. doi:10.1016/j.ijmultiphaseflow.2010.05.006

40. Chanson H (2011) Turbulent shear stresses in hydraulic jumps and decelerating surges: an experi-mental study. Earth Surf Process Landf 36(2):180–189. doi:10.1002/esp.2031

41. Chanson H (2012) Momentum considerations in hydraulic jumps and bores. J Irrig Drain Eng ASCE138(4):382–385. doi:10.1061/(ASCE)IR.1943-4774.0000409

42. Chanson H (2013) Hydraulics of aerated flows: Qui Pro Quo?. J Hydraul Res, IAHR. Invited Visionpaper, 51(3): 223–243. doi: 10.1080/00221686.2013.795917

43. Chanson H (2016) Atmospheric noise of a breaking tidal bore. J Acoust Soc Am 139(1):12–20. doi:10.1121/1.4939113

44. Chanson H, Cummings PD (1994) Effects of plunging breakers on the gas contents in the oceans. MarTechnol Soc J 28(3):22–32

72 Environ Fluid Mech (2017) 17:47–77

123

Page 27: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

45. Chanson H, Gualtieri C (2008) Similitude and scale effects of air entrainment in hydraulic jumps.J Hydraul Res 46(1):35–44

46. Chanson H, Montes JS (1995) Characteristics of undular hydraulic jumps. Experimental apparatus andflow patterns. J Hydraul Eng ASCE 121(2):129–144

47. Chanson H, Aoki S, Maruyama M (2002) Unsteady air bubble entrainment and detrainment at aplunging breaker: dominant time scales and similarity of water level variations. Coast Eng46(2):139–157

48. Chanson H, Aoki S, Hoque A (2006) Bubble entrainment and dispersion in plunging jet flows:freshwater versus seawater. J Coast Res 22(3):664–677. doi:10.2112/03-0112.1

49. Chanson H, Brattberg T (2000) Experimental study of the air–water shear flow in a hydraulic jump. IntJ Multiph Flow 26(4):583–607

50. Chanson H, Reungoat D, Simon B, Lubin P (2011) High-frequency turbulence and suspended sedi-ment concentration measurements in the Garonne River Tidal Bore. Estuar Coast Shelf Sci95(2–3):298–306. doi:10.1016/j.ecss.2011.09.012

51. Chanson H, Toi YH (2015) Physical modelling of breaking tidal bores: comparison with prototypedata. J Hydraul Res IAHR 53(2):264–273. doi:10.1080/00221686.2014.989458

52. Chen G, Kharif C, Zaleski S, Li J (1999) Two-dimensional Navier–Stokes simulation of breakingwaves. Phys Fluids 11(1):121–133

53. Cheng J, Wang P (2015) Extracting turbulence under breaking waves in the surf zone. J Waterw PortCoast Ocean Eng. doi:10.1061/(ASCE)WW.1943-5460.0000307

54. Christensen ED, Walstra D-J, Emerat N (2002) Vertical variation of the flow across the surf zone.Coast Eng 45:169–198

55. Cienfuegos R, Barthelemy E, Bonneton P (2010) Wave-breaking model for boussinesq-type equationsincluding roller effects in the mass conservation equation. J Waterw Port Coast Ocean Eng136(1):10–26

56. Cipriano RJ, Blanchard DC (1981) Bubble and aerosol spectra produced by a laboratory ‘breakingwave’. J Geophys Res 86(C9):8085–8092

57. Clift R, Grace JR, Weber ME (1978) Bubbles, drops, and particles. Academic Press, San Diego58. Coakley DB, Haldeman PM, Morgan DG, Nicolas KR, Penndorf DR, Wetzel LB, Weller CS (2001)

Electromagnetic scattering from large steady breaking waves. Exp Fluids 30(5):479–487. doi:10.1007/s003480000220

59. Cointe R, Tulin MP (1994) A theory of steady breakers. J Fluid Mech 216:1–2060. Cox DT, Shin S (2003) Laboratory measurements of void fraction and turbulence in the Bore Region

of surf zone waves. J Eng Mech ASCE 129(10):1197–120561. Crowe CT, Schwarzkopf JD, Sommerfeld M, Tsuji Y (2011) Multiphase flows with droplets and

particles, 2nd edn. Boca Raton, CRC Press62. D’Alessandro F, Tomasicchio GR (2008) The BCI criterion for the initiation of breaking process in

Boussinesq-type equations wave models. Coast Eng 55:1174–118463. Dabiri D, Gharib M (1997) Experimental investigation of the vorticity generation within a spilling

water wave. J Fluid Mech 330:113–13964. Dahl PH, Jessup AT (1995) On bubble clouds produced by breaking waves: an event analysis of ocean

acoustic measurements. J Geophys Res 100(C3):5007–5020. doi:10.1029/94JC0301965. Darcy HPG, Bazin H (1865) Recherches Hydrauliques. (Hydraulic Research.) Imprimerie Imperiales,

Paris, France, Parties 1ere et 2eme (in French)66. Deane G (1997) Sound generation and air entrainment by breaking waves in the surf zone. J Acoust

Soc Am 102(5):2671–2689. doi:10.1121/1.42032167. Deane G, Stokes MD (2002) Scale dependence of bubble creation mechanisms in breaking waves.

Nature 418:839–84468. Derakhti M, Kirby JT (2014) Bubble entrainment and liquid–bubble interaction under unsteady

breaking waves. J Fluid Mech 761:464–506. doi:10.1017/jfm.2014.63769. Docherty NJ, Chanson H (2012) Physical modelling of unsteady turbulence in breaking tidal bores.

J Hydraul Eng ASCE 138(5):412–419. doi:10.1061/(ASCE)HY.1943-7900.000054270. Drazen DA, Melville WK, Lenain L (2008) Inertial scaling of dissipation in unsteady breaking waves.

J Fluid Mech 611:307–33271. Drew DA (1983) Mathematical modeling of two-phase flow. Annu Rev Fluid Mech 1983(15):261–29172. Drew DA, Passman SL (1999). Theory of Multicomponent Fluids. Applied Mathematical Sciences,

Vol. 135, Springer: New York, doi:10.1007/b9767873. Duncan JH (1981) An experimental investigation of breaking waves produced by a towed hydrofoil.

Proc R Soc Lond A 377:331–34874. Duncan JH (2001) Spilling breakers. Annu Rev Fluid Mech 33:519–547

Environ Fluid Mech (2017) 17:47–77 73

123

Page 28: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

75. Esmaeeli A, Tryggvason G (1996) An inverse energy cascade in two-dimensional low Reynoldsnumber bubbly flows. J Fluid Mech 314:315–330

76. Falvey HT (1990) Cavitation in chutes and spillways. US Bureau of Reclamation EngineeringMonograph, No. 42, Denver, Colorado, USA

77. Farahani RJ, Dalrymple RA (2014) Three-dimensional reversed horseshoe vortex structures underbroken solitary waves. Coast Eng 91:261–279

78. Favre H (1935) Etude Theorique et Experimentale des Ondes de Translation dans les CanauxDecouverts. (‘Theoretical and Experimental Study of Travelling Surges in Open Channels.’) Dunod,Paris, France (in French)

79. Fuhrboter A (1970). Air entrainment and energy dissipation in breakers. 12th International Conferenceon Coastal Engineering. https://journals.tdl.org/icce/index.php/icce/article/view/2628

80. Furgerot L, Mouaze D, Tessier B, Perez L, Haquin S (2013) Suspended Sediment Concentration inRelation to the Passage of a Tidal Bore (See River Estuary, Mont Saint Michel, NW France). Proc.Coastal Dynamics 2013, Arcachon, France, 24–28 June, pp. 671–682

81. Galinat S, Risso F, Masbernat O, Guiraud P (2007) Dynamics of drop breakup in inhomogeneousturbulence at various volume fractions. J Fluid Mech 578:85–94. doi:10.1017/S0022112007005186

82. Gemmrich J (2010) Strong turbulence in the wave crest region. J Phys Oceanogr 40:583–595. doi:10.1175/2009JPO4179.1

83. Gemmrich JR, Farmer DM (2004) Near-surface turbulence in the presence of breaking waves. J PhysOceanogr 34:1067–1086

84. Gomez-Gesteira M, Rogers BD, Dalrymple RA, Crespo AJC (2010) State-of-the-art of classical SPHfor free-surface flows. J Hydraul Res 48(S1):6–27. doi:10.1080/00221686.2010.9641242

85. Govender K, Mocke GP, Alport MJ (2002) Video-imaged surf zone wave and roller structures andflow fields. J Geophys Res 107(C7):3072. doi:10.1029/2000JC000755

86. Gualtieri C, Chanson H (2010) Effect of Froude number on bubble clustering in a hydraulic jump.J Hydraul Res IAHR 48(4):504–508

87. Haller MC, Catalan PA (2010) Remote sensing of wave roller lengths in the laboratory. J Geophys Res114:C07022. doi:10.1029/2008JC005185

88. Henderson FM (1966) Open channel flow. MacMillan Company, New York89. Hinze JO (1955) Fundamentals of the hydrodynamic mechanism of slitting in dispersion processes.

J Am Inst Chem Eng 1:289–29590. Hoque A, Aoki S-I (2005) Distributions of void fraction under breaking waves in the surf zone. Ocean

Eng 32:1829–184091. Hornung HG, Willert C, Turner S (1995) The flow field downstream of a hydraulic jump. J Fluid Mech

287:299–31692. Hoyt JW, Sellin RHJ (1989) Hydraulic jump as ‘mixing layer’. J Hydraul Eng ASCE

115(12):1607–161493. Hubbard DW, Griffin OM, Peltzer RD (1987) Foam generation and air entrainment near a free surface.

Naval Research Laboratory Memorandum Report 6038, Washington D.C., USA94. Hwung HH, Chyan JM, Chung YC (1992) Energy dissipation and air bubbles mixing inside surf zone.

Proceedings of 23rd International Conference on Coastal Engineering, ASCE, Venice, Italy, Vol. 1,Chap. 22, pp. 308–321

95. Iafrati A (2009) Numerical study of the effects of the breaking intensity on wave breaking flows.J Fluid Mech 622:371–411

96. Iafrati A (2011) Energy dissipation mechanisms in wave breaking processes: spilling and highlyaerated plunging breaking events. J Geophys Res 116:C07024. doi:10.1029/2011JC007038

97. Jansen PCM (1986) Laboratory observations of the kinematics in the aerated region of breakingwaves. Coast Eng 9:453–477

98. Kalvoda PM, Xu L, Wu J (2003) Macrobubble clouds produced by breaking wind waves: a laboratorystudy. J Geophys Res 108(C6):3207. doi:10.1029/1999JC000265

99. Katz Y, Horev E, Wygnanski I (1992) The forced turbulent wall jet. J Fluid Mech 242:577–609100. Kazolea M, Delis AI, Synolakis CE (2014) Numerical treatment of wave breaking on unstructured

finite volume approximations for extended Boussinesq-type equations. J Comput Phys 271:281–305101. Kimmoun O, Branger H (2007) A particle image velocimetry investigation on laboratory surf-zone

breaking waves over a sloping beach. J Fluid Mech 588:353–397102. Kiger KT, Duncan JH (2012) Air-entrainment mechanisms in plunging jets and breaking waves. Ann

Rev Fluid Mech 44:563–596103. Kennedy AB, Chen Q, Kirby JT, Dalrymple RA (2000) Boussinesq modeling of wave transformation,

breaking and run-up. I: one dimension. ASCE. J Waterw Port Coast Ocean Eng 126(1):48–56

74 Environ Fluid Mech (2017) 17:47–77

123

Page 29: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

104. Koch C, Chanson H (2008) Turbulent mixing beneath an undular bore front. J Coast Res24(4):999–1007. doi:10.2112/06-0688.1

105. Koch C, Chanson H (2009) Turbulence measurements in positive surges and bores. J Hydraul ResIAHR 47(1):29–40. doi:10.3826/jhr.2009.2954

106. Kucukali S, Chanson H (2008) Turbulence measurements in hydraulic jumps with partially-developedinflow conditions. Exp Thermal Fluid Sci 33(1):41–53. doi:10.1016/j.expthermflusci.2008.06.012

107. Labourasse E, Lacanette D, Toutant A, Lubin P, Vincent S, Lebaigue O, Caltagirone J-P, Sagaut P(2007) Towards large eddy simulation of isothermal two-phase flows: governing equations and a prioritests. Int J Multiph Flow 33(1):1–39

108. Lakehal D, Liovic P (2011) Turbulence structure and interaction with steep breaking waves. J FluidMech 674:522–577. doi:10.1017/jfm.2011.3

109. Lamarre E, Melville WK (1991) Air entrainment and dissipation in breaking waves. Nature351:469–472. doi:10.1038/351469a0

110. Lamarre E, Melville WK (1992) Instrumentation for the measurement of void-fraction in breakingwaves: laboratory and field results. IEEE J Ocean Eng 17(2):204–215. doi:10.1109/48.126977

111. Lamarre E, Melville WK (1994) Void-fraction measurements and sound-speed fields in bubble plumesgenerated by breaking waves. J Acoust Soc Am 95(3):1317–1328. doi:10.1121/1.408572

112. Leandro J, Carvalho R, Chachereau Y, Chanson H (2012) Estimating void fraction in a hydraulic jumpby measurements of pixel intensity. Exp Fluids 52(5):1307–1318. doi:10.1007/s00348-011-1257-1

113. Leifer I, Caulliez G, de Leeuw G (2006) Bubbles generated from wind-steepened breaking waves: 2.Bubble plumes, bubbles, and wave characteristics. J Geophys Res 111:C06021. doi:10.1029/2004JC002676

114. Leng X, Chanson H (2015) Turbulent advances of a breaking bore: preliminary physical experiments.Exp Thermal Fluid Sci 62:70–77. doi:10.1016/j.expthermflusci.2014.12.002

115. Leng X, Chanson H (2015) Breaking bore: physical observations of roller characteristics. Mech ResCommun 65:24–29. doi:10.1016/j.mechrescom.2015.02.008

116. Leng X, Chanson H (2016) Coupling between free-surface fluctuations, velocity fluctuations andturbulent reynolds stresses during the upstream propagation of positive surges, bores and compressionwaves. Environ Fluid Mech 16(4):695–719. doi:10.1007/s10652-015-9438-8

117. Lennon JM, Hill DF (2006) Particle image velocimetry measurements of undular and hydraulic jumps.J Hydraul Eng 132(12):1283–1294

118. Lighthill J (1978) Waves in fluids. Cambridge University Press, Cambridge119. Lim H-J, Chang K-A, Huang Z-C, Na B (2015) Experimental study on plunging breaking waves in

deep water. J Geophys Res Oceans 120:2007–2049. doi:10.1002/2014JC010269120. Lin C, Hwung HH (1992) External and internal flow fields of plunging breakers. Exp Fluids

12:229–237121. Lin JC, Rockwell D (1995) Evolution of a quasi-steady breaking wave. J Fluid Mech 302:29–44122. Loewen MR, O’Dor MA, Skafel MG (1996) Bubbles entrained by mechanically generated breaking

waves. J Geophys Res: Oceans 101(C9):20759–20769123. Longuet-Higgins MS (1973) A model of flow separation at a free surface. J. Fluid Mech

57(1):129–148124. Longuet-Higgins MS, Turner JS (1974) An ‘entraining plume’ model of a spilling breaker. J. Fluid

Mech 63(1):1–20125. Lubin P, Vincent S, Abadie S, Caltagirone J-P (2006) Three-dimensional large eddy simulation of air

entrainment under plunging breaking waves. Coast Eng 53(8):631–655126. Lubin P, Glockner S (2015) Numerical simulations of three-dimensional plunging breaking waves:

generation and evolution of aerated vortex filaments. J Fluid Mech 767:364–393. doi:10.1017/jfm.2015.62

127. Ma G, Shi F, Kirby JT (2011) A polydisperse two-fluid model for surf zone bubble simulation.J Geophys Res 116:C05010

128. Madsen PA (1981) A model for a turbulent bore. Ph.D. Thesis, Tech. Univ. of Denmark, Inst. ofHydrodynamics and Hyd. Eng., Copenhagen, Denmark, 149 pages. (also Series Paper No. 28, Tech.Univ. of Denmark, Inst. of Hydrodyamics and Hyd. Eng., Copenhager, Denmark, 149 pages.)

129. Melville WK (1996) The role of surface-wave breaking in air–sea interaction. Annu Rev Fluid Mech28:279–321

130. Melville WK, Veron F, White CJ (2002) The velocity field under breaking waves: coherent structuresand turbulence. J Fluid Mech 454:203–233

131. Miller RL (1972) Study of air entrainment in breaking waves. Am Geophys Union Trans EOS 53:426132. Miller RL (1976) Role of vortices in surf zone prediction: sedimentation and wave forces. In Beach

and Nearshore Sedimentation, Edited by Richard A. Davis, Jr. and R. L. Ethington, 24: 92–114

Environ Fluid Mech (2017) 17:47–77 75

123

Page 30: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

133. Monahan EC (1993) Occurrence and evolution of acoustically relevant sub-surface bubble plumes andtheir associated, remotely monitorable, surface whitecaps’’. In: Kerman BR (ed) Natural physicalsources of underwater sound. Kluwer, Dordrecht, pp 503–517

134. Moraga FJ, Carrica PM, Drew DA, Lahey RT Jr (2008) A sub-grid air entrainment model for breakingbow waves and naval surface ships. Comput Fluids 37:281–298

135. Mori N, Suzuki T, Kakuno S (2007) Experimental study of air bubbles and turbulence characteristicsin the surf zone. J Geophys Res 112:C05014. doi:10.1029/2006JC003647

136. Mori N, Kakuno S (2008) Aeration and bubble measurements of coastal breaking waves. Fluid DynRes 40:616–626

137. Mossa M, Tolve U (1998) Flow visualization in bubbly two-phase hydraulic jump. J Fluids Eng,ASME 120:160–165

138. Mouaze D, Chanson H, Simon B (2010) Field measurements in the tidal bore of the Selune River inthe Bay of Mont Saint Michel (September 2010). Hydraulic Model Report No. CH81/10, School ofCivil Engineering, The University of Queensland, Brisbane, Australia

139. Murzyn F, Chanson H (2009) Free-surface fluctuations in hydraulic jumps: experimental observations.Exp Thermal Fluid Sci 33(7):1055–1064. doi:10.1016/j.expthermflusci.2009.06.003

140. Murzyn F, Mouaze D, Chaplin JR (2005) Optical fibre probe measurements of bubbly flow inhydraulic jumps. Intl J Multiph Flow 31(1):141–154

141. Musumeci RE, Svendsen IA, Veeramony J (2005) The flow in the surf zone: a fully nonlinearBoussinesq-type of approach. Coast Eng 52:565–598. doi:10.1016/j.coastaleng.2005.02.007

142. Nadaoka K, Hino M, Koyano Y (1989) Structure of the turbulent flow field under breaking waves inthe surf zone. J Fluid Mech 204:359–387

143. Oguz HN, Prosperetti A, Kolaini AR (1995) Air entrapment by a falling mass of water. J Fluid Mech294:181–207

144. Ohtsu, I., Yasuda, Y., and Gotoh, H. (2001). Hydraulic condition for undular-jump formations.J Hydraul Res IAHR, 39(2), 203–209. Discussion: 2002 40 (3), 379–384

145. Okamoto T, Basco DR (2006) The relative trough Froude number for initiation of wave breaking:theory experiments and numerical model confirmation. Coast Eng 53:675–690

146. Peregrine DH (1983) Breaking waves on beaches. Annu Rev Fluid Mech 15:149–178147. Peregrine DH, Svendsen IA (1978) Spilling breakers, bores and hydraulic jumps. Proc Int Conf Coast

Eng 30:540–550148. Prosperetti A (1988) Bubble-related ambient noise in the ocean. J Acoust Soc Am 84(3):1042–1054149. Rajaratnam N (1962) An experimental study of air entrainment characteristics of the hydraulic jump.

J Inst Eng India 42(7):247–273150. Rajaratnam, N. (1965). The Hydraulic Jump as a Wall Jet. Jl of Hyd. Div., ASCE, Vol. 91, No. HY5,

pp. 107–132. Discussion: Vol. 92, No. HY3, pp. 110–123 & Vol. 93, No. HY1, pp. 74–76151. Rao NSL, Kobus HE (1971) Characteristics of self-aerated free-surface flows. Water and waste water/

current research and practice, vol 10. Eric Schmidt Verlag, Berlin152. Rapp RJ, Melville WK (1990) Laboratory measurements of deep-water breaking waves. Philos Trans

R Soc Lond Ser A 331:735–800. doi:10.1098/rsta.1990.0098153. Resch FJ, Leutheusser HJ (1972) Le Ressaut Hydraulique: mesure de Turbulence dans la Region

Diphasique. Jl La Houille Blanche 1(4):279–293 (in French)154. Reungoat D, Chanson H, Caplain B (2014) Sediment processes and flow reversal in the undular tidal

bore of the Garonne River (France). Environ Fluid Mech 14(3):591–616. doi:10.1007/s10652-013-9319-y

155. Reungoat D, Chanson H, Keevil CE (2015) Field Measurements of Unsteady Turbulence in a TidalBore: the Garonne River in October 2013. J Hydraul Res IAHR. doi:10.1080/00221686.2015.1021717

156. Rojas G, Loewen MR (2010) Void fraction measurements beneath plunging and spilling breakingwaves. J Geophys Res 115:C08001. doi:10.1029/2009JC005614

157. Rouse H, Siao TT, Nagaratnam S (1959) Turbulence characteristics of the hydraulic jump. TransASCE 124:926–950

158. Ryabenko, A.A. (1990). Conditions favorable to the existence of an undulating jump.Gidrotekhnicheskoe Stroitel’stvo, No. 12, pp. 29–34 (in Russian). (Translated in HydrotechnicalConstruction, 1990, Plenum Publ., pp. 762–770)

159. Russell SO, Sheehan GJ (1974) Effect of entrained air on cavitation damage. Can J Civil Eng1:97–107

160. Salter ME, Nilsson ED, Butcher A, Bilde M (2014) On the seawater temperature dependence of the seaspray aerosol generated by a continuous plunging jet. J Geophys Res Atmos. doi:10.1002/2013JD021376

76 Environ Fluid Mech (2017) 17:47–77

123

Page 31: Are breaking waves, bores, surges and jumps the same flow?staff.civil.uq.edu.au/h.chanson/reprints/Lubin_Chanson_efmc_2017.pdf · Keywords Breaking waves Breaking bores Hydraulic

161. Schaffer HA, Madsen PA, Deigaard R (1993) A Boussinesq model for waves breaking in shallowwater. Coast Eng 20:185–202

162. Shi F, Kirby JT, Ma G (2010) Modeling quiescent phase transport of air bubbles induced by breakingwaves. Ocean Model 35:105–117

163. Simpson JH, Fisher NR, Wiles P (2004) Reynolds Stress and TKE Production in an Estuary with atidal bore. Estuar Coast Shelf Sci 60(4):619–627

164. Stoker JJ (1957) Water waves. The mathematical theory with applications. New York, IntersciencePublishers

165. Stokes MD, Deane GB (1999) A new optical instrument for the study of bubbles at high void fractionswithin breaking waves. IEEE J Ocean Eng 24(3):300–311

166. Stokes D, Deane G, Vagle S, Farmer D (2002) Measurements of large bubbles in open-oceanwhitecaps. Gas Transf Water Surf, pp. 279–284. doi:10.1029/GM127p0279

167. Svendsen IA (1984) Wave heights and set-up in a surf zone. Coast Eng 8:303–329168. Svendsen IA (1984) Mass flux and undertow in a surf zone. Coast Eng 8:347–365169. Taylor G (1934) The formation of emulsions in definable field of flow. Proc R Soc A 146:501–523170. Thorpe SA (1982) On the clouds of bubbles formed by breaking wind-waves in deep water, and their

role in air–sea gas transfer. Philos Trans R Soc Lond A 304:155–210. doi:10.1098/rsta.1982.0011171. Tissier M, Bonneton P, Marche F, Lannes D (2012) A new approach to handle wave breaking in fully

non-linear Boussinesq models. Coast Eng 67:54–66172. Tonelli M, Petti M (2009) Hybrid finite volume-finite difference scheme for 2DH improved Boussi-

nesq equations. Coast Eng 56:609–620173. Treske A. (1994). Undular bores (favre-waves) in open channels—experimental studies. J Hydraul

Res, IAHR, Vol. 32, No. 3, pp. 355–370. Discussion: Vol. 33, No. 3, pp. 274–278174. Tricker RAR (1965) Bores, breakers waves and wakes. American Elsevier Publ. Co., New York175. Vagle S, Farmer DM (1998) A comparison of four methods for bubble size and void fraction mea-

surements. IEEE J Ocean Eng 23(3):211–222. doi:10.1109/48.701193176. Veron F (2015) Ocean spray. Annu Rev Fluid Mech 47:507–538. doi:10.1146/annurev-fluid-010814-

014651177. Violeau D, Rogers BD (2016) Smoothed particle hydrodynamics (SPH) for free-surface flows: past,

present and future. J Hydraul Res. doi:10.1080/00221686.2015.1119209178. Wang H, Chanson H (2015) Air entrainment and turbulent fluctuations in hydraulic jumps. Urban

Water J 12(6):502–518. doi:10.1080/1573062X.2013.847464179. Wang H, Murzyn F, Chanson H (2014) Total pressure fluctuations and two-phase flow turbulence in

hydraulic jumps. Exp Fluids 55(11):1–16. doi:10.1007/s00348-014-1847-9180. Wang H, Murzyn F, Chanson H (2015) Interaction between free-surface, two-phase flow and total

pressure in hydraulic jump. Exp Thermal Fluid Sci 64:30–41. doi:10.1016/j.expthermflusci.2015.02.003

181. Wang P, Yao Y, Tulin MP (1995) An efficient numerical tank for non-linear water waves, based on themulti-subdomain approach with BEM. Int J Numer Methods Fluids 20(12):1315–1336

182. Wanninkhof R, Asher WE, Ho DT, Sweeney C, McGillis WR (2009) Advances in quantifying air–seagas exchange and environmental forcing. Annu Rev Fluid Mech 1:213–244

183. Wolanski E, Williams D, Spagnol S, Chanson H (2004) Undular tidal bore dynamics in the DalyEstuary, Northern Australia. Estuar Coast Shelf Sci 60(4):629–636

184. Wood IR (1991) Air entrainment in free-surface flows. IAHR hydraulic structures design manual No.4, hydraulic design considerations. Balkema Publ, Rotterdam

185. Zhang G, Wang H, Chanson H (2013) Turbulence and aeration in hydraulic jumps: free-surfacefluctuation and integral turbulent scale measurements. Environ Fluid Mech 13(2):189–204. doi:10.1007/s10652-012-9254-3

Environ Fluid Mech (2017) 17:47–77 77

123